You are on page 1of 18

R e s e a rc h

A review on biomass torrefaction process


and product properties for energy
applications
Jaya Shankar Tumuluru,1* Shahab Sokhansanj,2 J. Richard Hess,1 Christopher T. Wright,1 and Richard D. Boardman1

Abstract
Torrefaction of biomass can be described as a mild form of pyrolysis at temperatures typically ranging between 200 and 300°C in an inert and
reduced environment. Common biomass reactions during torrefaction include devolatilization, depolymerization, and carbonization of hemicellulose,
lignin, and cellulose. The torrefaction process produces a brown to black uniform solid product, as well as condensable (water, organics, and lipids) and
­noncondensable gases (CO2, CO, and CH4). Typically during torrefaction, 70% of the mass is retained as a solid product, containing 90% of the initial
energy content, while 30% of the lost mass is converted into condensable and noncondensable products. The system’s energy efficiency can be improved
by reintroducing the material lost during torrefaction as a source of heat. Torrefaction of biomass improves its physical properties like grindability; par-
ticle shape, size, and distribution; pelletability; and proximate and ultimate composition like moisture, carbon and hydrogen content, and calorific value.
Compared to raw biomass, the carbon content and calorific value of torrefied biomass increases by 15–25% wt, while the moisture content decreases to
<3% (w.b.). Torrefaction decreases the grinding energy by about 70%, and the ground torrefied biomass has improved sphericity, particle surface area, and
particle size distribution. Torrefied biomass pelletization at temperatures of 225°C decreases the specific energy consumption and increases the capacity
of the mill by a factor of 2. The loss of the OH functional group during torrefaction makes the material hydrophobic (i.e., loses the ability to attract water
molecules) and more stable against chemical oxidation and microbial degradation. These improved properties make torrefied biomass particularly suitable
for cofiring in power plants and as an upgraded feedstock for gasification.

Idaho National Laboratory, 2525 Fremont Avenue, Idaho Falls, Idaho 83415,
1
from different thermochemical (combustion, gasification, and pyrolysis),
USA 2Oak Ridge National Laboratory, PO Box 2008, Oak Ridge, Tennesse biological (anaerobic digestion and ­fermentation), or chemical (esteri-
37831, USA
fication) processes, where direct combustion can provide a near-term
*Author for correspondence. Email: jayashankar.tumuluru@inl.gov energy solution.1 The growing interest in biomass as a solid fuel includes
Submitted: 5 August 2011; Revised: 16 September 2011; Accepted: 17 combustion to produce steam for electrical power and commercial plant
September 2011 uses, as well as gasification to produce a combustible gas (large partial
DOI: 10.1089/ind.2011.0014 pressure of nitrogen and CO2, called producer gas) and syngas (carbon
monoxide and hydrogen with low amounts of nitrogen and CO2). Still,
Keywords: biomass torrefaction; torrefaction temperature and residence the use of either producer gas or syngas in modern reciprocating or gas
time; biomass reactions; off-gas composition; torrefied biomass physical,
chemical, and storage properties turbines, or to produce higher value chemicals and fuels, is limited due to
biomass feedstock preparation, accumulation logistics, and economics.
ABBREVIATIONS: CV, calorific value; GHG, greenhouse gas; TOP, torrefaction
and pelletization Limitations of biomass as fuel
Some of the inherent problems with raw biomass materials
Introduction

T
c­ ompared to fossil fuel resources (low bulk density, high moisture
he world is currently challenged to reduce dependence on fos- content, hydrophilic nature, and low calorific value [CV])  render
sil fuels and achieve a sustainable, renewable ­energy supply. raw biomass difficult to use on a large scale.1 These limitations
Energy produced from biomass is considered ­carbon-neutral greatly impact logistics and final energy efficiency. Due to its
because the carbon dioxide released during conversion is low energy density compared to fossil fuels, very high volumes of
already part of the carbon cycle.1 Increasing biomass use for energy can biomass are needed, which compounds problems associated with
help reduce greenhouse gas (GHG) emissions and meet the targets estab- storage, transportation, and feed handling at cogeneration, thermo-
lished in the Kyoto Protocol.2 Energy from biomass can be produced chemical, and biochemical conversion plants.

384 INdustrial Biotechnology october 2011


research

High moisture in raw biomass is one of the primary challenges, shows how the pretreatment of biomass makes the biomass easier
as it reduces the efficiency of the process and increases fuel pro- to convert.5
duction costs.3 High moisture content in biomass leads to natural Torrefaction, which is a thermal pretreatment process, is a viable
decomposition, resulting in loss of quality and storage issues such technology that significantly alters the physical and chemical com-
as off-gas emissions. Another consequence of high moisture ­content position of the biomass. Torrefaction is defined as slowly heating bio-
is the uncertainty it causes in biomass’s physical, chemical, and mass in an inert environment and temperature range of ­200–300°C.
microbiological properties. Irregular biomass shapes constitute This  process improves the physical, chemical, and biochemical
another issue, especially during feeding in a cofiring or gasification composition of the biomass, making it perform better for cofiring
system. In addition, biomass has more oxygen than carbon and and gasification purposes. Many researchers have studied the effect
hydrogen, making it less suitable for thermochemical conversion of torrefaction process time and temperature on the physical and
processes. Considered collectively, these properties make raw bio- chemical composition.6–15 However, a detailed literature review is
mass ­unacceptable for energy applications. lacking on the torrefaction process in terms of biomass reactions
To overcome these challenges and make biomass suitable for (such as depolymerization, devolatilization, and carbonization) and
energy applications, the material must be preprocessed. One of the product properties. The primary focus of this research is to conduct
commonly used preprocessing operations is grinding, which helps to a detailed literature review on biomass ­torrefaction, which includes
achieve a consistent particle size; however, the moisture content of (a) biomass reactions, including chemical and structural changes, (b)
the biomass limits the performance of many grinders.3 Furthermore, torrefaction product yields in terms of condensable, noncondens-
grinding can be very costly when smaller particle sizes are desired able, and solid product, and (c) the ­solid torrefied product’s physical,
and, in some cases, impractical for biomass with high moisture con- chemical, and storage properties for energy applications.
tent. High moisture content can also result in inconsistent particle
sizes (especially when the particles are less than 2 mm), which may Biomass components
not react consistently, thereby reducing the efficiency and increasing The plant cell wall is the tough, usually flexible but sometimes
the costs of the conversion process. Also, raw biomass is thermally fairly rigid layer that provides structural support and protection
unstable due to high moisture, which results in low CVs and incon- from mechanical and thermal stresses. The major components of
sistent particle-size distribution issues when used in thermochemical the primary cell wall are cellulose (carbohydrates), hemicelluloses,
processes such as gasification. This can lead to inconsistent products and pectin. The cellulose microfibrils are linked via hemicellu-
and the formation of condensable tars, which results in problems like losic ­tethers to form the cellulose-hemicellulose network, which is
gas-line blockage.4 embedded in the pectin matrix. The outer part of the primary cell
A viable option is to pretreat the biomass before the end-use wall is usually impregnated with cutin and wax, forming a perme-
application. Pretreatment helps alter biomass’s physical properties ability barrier known as the plant cuticle.16
and chemical composition and makes it more suitable for conver- Cells and tissues of the plant body play an important role in the
sion.5 The pretreatment can be a chemical, thermal, or mechanical growth of the plant. The structural complexity of the plant body
process, like ammonia fiber explosion, torrefaction, and steam results from variations in the form and function of the cells and also
explosion, respectively. These pretreatment processes help alter the from differences in the manner of combination of cells into tissue
amorphous and crystalline regions of the biomass and bring sig- and a tissue system. The three different types of plant tissues are
nificant changes in structural and chemical compositions. Figure 1 (1) meristematic tissue, (2) dermal tissue, and (3) vascular tissue.17
Secondary cell walls contain a wide range of additional ­compounds
Cellulose that modify their mechanical properties and ­p ermeability. The
­polymers that make up the secondary cell wall include (1) ­cellulose,
Lignin
(2) xylan, a type of hemicellulose, (3) lignin, a complex phenolic
polymer that penetrates the spaces in the cell wall between cel-
lulose, hemicellulose, and pectin components and which drives out
Amorphous Pretreatment water and strengthens the wall, and (4) structural proteins (approxi-
mately 1–5%), which are found in most plant cell walls.16 Figure 2
region
shows the plant cell wall and lignocellulosic biomass composition.5
Table 1 shows the typical lignocellulosic content of some plant and
woody biomass.12,18
Crystalline
region Torrefaction process overview
Torrefaction is a thermal pretreatment technology. It is also defined
Hemicellulose
as isothermal pyrolysis of biomass occurring in temperature ranges of
Figure 1. Pretreatment effect on lignocellulosic biomass 5 200–300°C and performed at atmospheric pressure in the absence of

© mary ann liebert, inc. • Vol. 7 No. 5 • october 2011 INdustrial Biotechnology 385
research

Plant cell wall

Cellulose

Lignin

Hemicellulose

Figure 2. Plant cell wall and lignocellulosic biomass composition5

reaction is more exothermal, and the degradation of hemicellulose


Table 1. Typical lignocellulosic content of some plant
continues. At these temperatures, biomass begins to brown and give
materials12,18
off additional moisture, CO2, and large amounts of acetic acid with
Lignocellulosic content (%) some phenols that have low energy values.7
Plant material Hemicellulose Cellulose Lignin During torrefaction, the major decomposition reactions affect the
hemicellulose. Lignin and cellulose also decompose in the range of
Orchard grass
40.0 32.0 4.7 temperatures at which torrefaction is normally carried out, but to a
(medium maturity)
lesser degree.19–20 Torrefied biomass retains most of its energy and
Rice straw 27.2 34.0 14.2 simultaneously loses its hygroscopic properties. At about 280°C,
Birch wood 25.7 40.0 15.7 the reaction is entirely exothermic, and gas production increases,
resulting in the formation of carbon monoxide, hydrocarbons like
Reed canary grass 29.7 42.6 7.6
phenols and cresols, and other heavier products. For torrefaction,
Wheat straw 30.8 41.3 7.7 process temperatures over 300°C are not recommended, as these
Willow 14.1 49.3 20.0 may lead to extensive devolatilization of the biomass due to the
initiation of the pyrolysis process.8

Torrefaction process technique


oxygen.6–7 Biomass torrefaction has been recognized as a technically Torrefaction is a thermochemical process involving the interac-
feasible method of converting raw biomass into a solid that is suitable tion of drying and incomplete pyrolysis. The different parameters
for commercial and residential combustion and gasification applica- that influence the torrefaction process are (a) reaction temperature,
tions, given that it has high energy density, is hydrophobic, com- (b) heating rate, (c) absence of oxygen, (d) residence time, (e) ­ambient
pactable, and grindable, and has a lower oxygen-to-carbon (O/C) ratio. pressure, (d) flexible feedstock, (e) feedstock moisture, and (f) feed-
During the initial heating of lignocellulosic materials, water due stock particle size. Biomass feedstock is typically predried to 10%
to chemical reactions is removed through a thermocondensation or less moisture content prior to torrefaction. Particle size plays
process. This happens at temperatures between 160 and 180°C and an important role in torrefaction in that it influences the reaction
results in the formation of CO2.7 At temperatures of 180–270°C, the mechanisms, kinetics, and duration of the process, given a specific

386 INdustrial Biotechnology october 2011


research

heating rate. The chemical reactions that occur when reactive inter- dation depends on the duration of the heating and temperature.21–23
mediates are trapped in a thick matrix differ from the situations The various components in the biomass perform unique functions
in which products can escape and be swept away in a gas stream. within the plant structure and interact differently depending on
The duration of the process is basically adjusted to produce friable, thermal treatment time and temperature.
hydrophobic, and energy-rich enhanced biomass fuel. Drying process variables that can influence the structural and
chemical compositional changes include biomass composition,
Temperature effect on the biomass components particle size, processing temperature and time, heating rate, gas
Understanding plant components and their composition will composition, pressure, and flow rate.9 Figure 3 provides a modified
help in understanding the biomass degradation reactions during version of Bergman et al’s description of the torrefaction process.10
the thermal pretreatment process. Exposure of biomass to elevated The updated figure describes the structural changes, emissions due
temperatures results in thermal degradation of its structure, which to bond cleavage, biomass reactions, and color changes at different
is often accompanied by loss of mass. The degree of thermal degra- temperature regimes ranging from 50–300°C.9,21–30

Reactive drying (initiates


Nonreactive drying (no changes changes in chemical Destructive drying (alters
in chemical composition) composition) chemical composition)
Mostly surface Insignificant organic Initiation of hydrogen Breakage of inter- and intramolecular hydrogen,
Water, organic emissions,

moisture removal emissions and carbon bonds C-O and C-C bonds. Emission of hydrophilic
breaking. Emission extractives (organic liquid product having
of lipophylic compounds oxygenated compounds). Formation of higher
and gases

like saturated and molecular mass carboxylic acids


unsaturated fatty acids, (CH3-(CH2)n-COOH), n=10-30), alcohols,
sterols, terpenes, which aldehydes, ether and gases like CO, CO2 and CH4
have no capacity to form
hydrogen bonds.

Initial disruption of Maximum cell Structural Complete destruction of cell structure. Biomass
Cell and

cell structure structure disruption deformity loses its fibrous nature and acts very brittle.
tissue

and reduced
porosity
Hemicellulose

Extensive
Depolymerization Limited
devolatilization
Drying (A) and devolatilization and
recondensation and carbonization
(C) carbonization (D) (E)
Lignin

Glass transition/
A C D E
softening (B)
Cellulose

A C D E
Color changes
in biomass

Torrefaction

50 100 150 200 250 300


Temperature (°C)
Figure 3. Structural, chemical, and color changes in biomass at different drying temperatures (modified version of Bergman et al’s ­description
of the torrefaction process10)

© mary ann liebert, inc. • Vol. 7 No. 5 • october 2011 INdustrial Biotechnology 387
research

Based on Figure 3, at drying temperatures of 50–150°C (A), ­biomass different temperature regimes indicate that the biomass turns brown
loses moisture and shrinks. This also results in reduced porosity in the to black at 150–300°C, which can be mainly attributed to chemical
biomass, though the biomass may still have the ability to retain its compositional changes. Lam et al quantified the severity of steam
structure if rewetted. This region is known as the nonreactive drying treatment based on color coordinates (such as L, a, and b values,
zone, where most of the chemical constituents of the biomass remain which indicate lightness, redness/greenness, and yellowness/blue-
intact. At the higher end of these temperatures (i.e., 120–150°C) ness, respectively) in their study.33 The same authors also developed
(B), the lignin softens and makes the material more suitable for multilinear regression models to describe the chemical composi-
densification, as the softened lignin acts as a binder. Temperature tional changes like carbon and hydrogen based on color changes in
Regime C (i.e., 150–200°C), also called the reactive drying range, steam-exploded wood pellets.33 In addition, color measurement can
initiates the breakage of hydrogen and carbon bonds and results in be a good indicator for identifying impurities like bark, ash, or any
the emission of lipophilic extractives and compounds due to thermal other foreign material present in the biomass. For example, in the
degradation of biomass solids. This temperature also results in struc- coffee bean roasting process, the change in the color is used as an
tural ­deformity, from which biomass loses its ability to regain its indicator to define the changes in chemical composition.34 Typical
original structure if ­rewetted. Also, according to Bergman and Kiel, thermal treatment process variables (mass and energy yields at dif-
­depolymerization of ­hemicellulose results in shortened, condensed ferent temperature regimes) are given in Table 2.
polymers with solid structures.10,31 Increasing the temperature further, During torrefaction at 200–300°C, mass loss predominantly
as shown in Regime D, also called destructive drying (200–300°C), results from the loss of moisture and decomposition (devolatiliza-
results in carbonization and devolatilization. tion), particularly hemicellulose and some lignin. Xylan-based
These temperatures represent the torrefaction process limits, hemicellulose generally decomposes around 250–280°C. Lignin
which result in the disruption of most inter- and intramolecular decomposition proceeds more slowly, but gradually increases start-
hydrogen bonds and C–C and C–O bonds, resulting in the formation ing at about 200°C.11 However, the thermal decomposition behavior
of hydrophilic extractives, carboxylic acids, alcohols, aldehydes, of individual biomass polymers may be different from the strongly
ether, and gases like CO, CO2, and CH4. At these temperatures, cell integrated structure of whole biomass. Figure 4 indicates the typical
structure is completely destroyed as the biomass loses its fibrous weight loss in cottonwood as a function of temperature.35 The figure
nature and becomes brittle. Bergman reports further that increas- clearly shows that the weight loss and degradation of biochemical
ing the temperature to >300°C results in extensive devolatilization composition typically starts at temperatures higher than 200°C.
and carbonization of the polymers.10,32 The blue line in Figure 3
indicates that at temperatures lower than 250°C, the mass loss is at Biomass reactions
a ­minimum, as main biomass decomposition results from limited Cell and tissue: Changes in biomass cells and tissues, which lead
devolatilization and carbonization of the hemicellulose. At tem- to structural changes, typically happen at drying temperatures of
peratures >250°C, the hemicellulose decomposes extensively into 50–150°C. Terziev observed that different thermal treatments have a
volatiles and a char-like solid product. Lignin and cellulose show distinct effect on wood microstructure and properties.36 Some studies
limited devolatilization and carbonization. on drying wood in a 100–150°C temperature range indicated signifi-
Color change in the raw material can be a good means of describ- cant cell wall shrinkage and a pore size volume decrease, which is
ing the degree of torrrefaction. The color changes in the biomass at attributed to moisture loss.37 The drying damage manifests as irregu-

Table 2. Thermal treatment process variables for different temperature regimes


TEMPERATURE Time Process reactions Heating rate Drying Mass yield Energy yield
(°C) (min) (°C/min) environment (%) (%)
and pressure

50–150 30–120 Nonreactive drying (moisture removal and structural <50 Air and ambient ~90–95 Not
changes) pressure significant

150–200 30–120 Reactive drying (moisture removal and structural <50 Air and ambient ~90 Needs to
damage due to cell wall collapse) pressure be researched

200–300 <30 Destructive drying <50 Inert ~70 ~90


Ø Devolatilization and carbonization of hemicellulose environment
Ø Depolymerization and devolatilization/softening and ambient
of lignin pressure
Ø Depolymerization and devolatilization of cellulose

388 INdustrial Biotechnology october 2011


research

larly distributed micro-cracks within the biomass cell wall. Three c­ ondensable and noncondensable products. Thermal degradation
mechanisms result in structural changes to the biomass due to drying. of hemicellulose initiates at 150°C, with the majority of weight loss
First, initially elevated temperatures induce thermal degradation of the occurring above 200°C, depending on the chemical nature of the
biomass components, which results in formation of cavities within the hemicellulose and the relationship with lignin within the cell.18,49
cell.21–22,38 Second, the pore wall starts to collapse because water has Compared to cellulose, hemicellulose generally evolves as light
been removed. This will result in closure of the pore wall due to irre- volatiles, producing fewer tars and less char. Many researchers have
versible hydrogen bonding.39–40 Third, the anisotropic drying shrink- noted that major hemicellulose decomposition reactions occur at
age of cell wall layers results in drying stresses, which can damage temperatures between 220°C and 280°C.9 Native hemicellulose is
wood’s cell walls.41–43 partly depolymerized by hydrolysis and/or thermal chain scission to
Kauman explained that total cell wall collapse during drying provide “reacting” hemicellulose. This intermediate is decomposed
can be due to liquid tension collapse and stress collapse, which by acid and radical reactions to yield many substances (e.g., furfu-
lead to biomass shrinkage.44 It has been reported that the collapse ral) that recombine to form torrefied hemicellulose. Water and acids
intensity increases with increased temperature during drying. 45–48 form during the above reactions and are released into the reaction
Borrega  and Kärenlampi stated that exposure of wood to tem- environment.50 Some of this water may be reused to depolymerize
peratures of >150°C during drying may cause thermal degradation hemicellulose or to release acids from the hemicellulose by hydro-
of its structure, often accompanied by loss of mass.38 The degree lysis of acetate groups. These radical reactions can also result in the
of structural damage depends on tree species as well as process formation of acids.
parameters, such as duration of treatment, temperature, and relative
humidity.21–23 All structural changes observed in the biomass due Cellulose: Cellulose is a high-molecular-weight polymer (molecular
to moisture loss influence the biomass’s mechanical properties, like weight of 106 and higher) that makes up the fibers in wood and other
bulk density and grindability. biomass. The multiple hydroxyl groups on the glucose residues from
one chain form hydrogen bonds with oxygen molecules on the same
Hemicellulose: Hemicellulose is a branched polymer (also a poly- or a neighboring chain. These chains are firmly held together side
saccharide) that consists of shorter chains (500–3000 sugar units) by side, resulting in the formation of microfibrils with high tensile
as opposed to the 7000–15 000 glucose molecules per unbranched strength. This strength is important in cell walls, where the micro-
polymer observed in cellulose, as shown in Figure 2. Hemicellulose fibrils are meshed into a carbohydrate matrix, conferring rigidity to
accounts for about 25–35 wt% of biomass. The polysugars that plant cells. Cellulose strands with hydrogen bonds (dashed within
form hemicelluloses vary differently among woods and much more and between cellulose molecules) are shown in Figure 2.
among plants. In the case of deciduous wood, the hemicellulose Cellulose degradation occurs between 240 and 350°C, resulting
fraction is very different. Deciduous wood contains 80–90% of in anhydrous cellulose and levoglucosan.18 The crystalline struc-
4-O-methylglucoronoxylan (referred to as xylan), whereas conifer- ture resists thermal depolymerization better than do unstructured
ous wood (most softwood) contains 60–70% of glucomannan and hemicelluloses. Amorphous regions in the cellulose contain waters
15–30 wt% of arabinogalactan.18 of hydration and hold free water within the plant. When heated
Hemicellulose undergoes major decomposition reactions at rapidly, this water is converted to steam, which can further rupture
­t orrefaction temperatures of 200–300°C, resulting in different the cellulose structure.
Cellulose also undergoes polymer restructuring similar to hemi-
100 cellulose through depolymerization reactions, but to a lesser extent.
90 The water and acids released due to hemicellulose depolymerization
Acid lignin
80 can result in more cellulose degradation. This disordered cellulose is
thermostable and contains furan, aliphatic, and keto groups.
70
Weight (%)

60 Milled
wood lignin Lignin: Lignin is an amorphous, highly branched, cross-linked
50 macromolecular polyphenolic resin with no exact structure. Lignin
40 fills the spaces in the cell wall between cellulose, hemicellulose,
Cellulose and pectin components. It is covalently linked to hemicellulose
30 Xylan
Wood and thereby cross-links different plant polysaccharides, conferring
20
mechanical strength to the cell wall and, by extension, to the plant
10 as a whole. It is relatively hydrophobic and aromatic in nature. The
0 degree of polymerization in nature is difficult to measure since it is
100 200 300 400 500 fragmented during extraction and the molecule consists of various
Temperature (°C) types of substructures that appear to repeat in a haphazard manner.
Figure 4. Thermogravimetry of cottonwood and its constituents35 An example of a lignin polymer is provided in Figure 2.

© mary ann liebert, inc. • Vol. 7 No. 5 • october 2011 INdustrial Biotechnology 389
research

Lignin decomposes when heated to 280–500°C, producing Torrefaction products


­ henols via the cleavage of ether bonds and scissioning of ­carbon–
p During torrefaction, three different products are produced:
carbon bonds.18,49 Lignin is difficult to dehydrate and thus converts (1) brown to black uniform solid biomass, which is used for bioen-
to char more than cellulose or hemicellulose. Evans et al showed ergy applications, (2) condensable volatile organic compounds com-
that lignin releases some small phenolic fragments within the prising water, acetic acid, aldehydes, alcohols, and ketones, and (3)
­torrefaction temperature range.51 Grafting torrefied hemicellulose noncondensable gases like CO2, CO, and small amounts of methane.
onto cellulose and/or lignin may also be possible, indicating that Release of these condensable and noncondensable products results
all of the biomass polymers can participate in torrefaction.9 The in changes in the physical, chemical, and storage properties of
biomass decomposition reactions at different temperature regimes biomass. Several studies have also investigated the physical prop-
are given in Figure 5.28 erties and chemical composition of the liquids and gases released
during torrefaction.11,14,52 Bergman et al provide an overview of the
torrefaction products, based on their states at room temperature,
char CO2+CO+H2 <250°C which can be solid, liquid, or gas, as shown in Figure 6.10 The solid
Biomass phase consists of a chaotic structure of the original sugars and
char+tar+CO2+CO+H20+H2+CH2+CnHm >400°C reaction products. The gas phase includes gases that are considered
H2+CH2+CnHm >700°C+residence time permanent gases, and light aromatic components such as benzene
and toluene.
Figure 5. Biomass decomposition reactions at different temperature The condensables, or liquids, can be further divided into four
regimes28 subgroups: (1)  reaction water produced from thermal decom-
position, (2) freely bound water that has been released through
evaporation, (3) organics (in liquid form), which consist of organics
Organic extractables: Organic extractables include fats, waxes, produced during devolatilization and carbonization, and (4) lipids,
alkaloids, proteins, phenolics, simple sugars, pectins, mucilages, which contain compounds such as waxes and fatty acids.
gums, resins, terpenes, starches, glycosides, saponins, and essential Condensable and noncondensable products are emitted from
oils.18 Again, these vary in amount and composition depending on the biomass based on heating rate, torrefaction temperature and
the type of biomass. Hence, devolatilization behavior depends on time, and biomass composition. The emission profiles of these
the amount and location of the biomass extractables. products greatly depend on the moisture content in the biomass.

After torrefaction
Phase Groups of components
• Original sugar structures
• Modified sugar structures
Solid • Newly formed polymeric structures
• Char
• Ash

H2O
Biomass Organics:
Liquid (condensable) Sugars, polysugars, acids, alcohols, furans, ketones
Lipids
Terpenes, phenols, fatty acids, waxes, tanins

• H2, CO, CO2, CH4


Gas (permanent)
• CxHy, toluene, benzene

Figure 6. Products formed during torrefaction of biomass10

390 INdustrial Biotechnology october 2011


research

At high moisture content, some of the surfactants present in the Comparing the conversion of agricultural residues (i.e., rice straw
biomass (oleic acid, fatty acids, and resin acids) help solubilize and rape stalk) to woody biomass from the literature, the solid-to-
the ­m onoterpenes in water and transport them through the liquid conversion of the former is much higher than that of the
wood matrix. At low moisture content, the emission of different latter under the same torrefaction conditions (i.e., temperature and
compounds like terpenes is driven by vapor pressure and tem- residence time).10 This is due to the higher volatile matter content
perature. During ­torrefaction the raw material loses most of its in the agricultural residues and the decomposition of hemicellu-
moisture and other volatiles, which have a low heat value. Many lose, the main fraction decomposed in the torrefaction temperature
researchers have worked on identifying gas composition in terms range. Bridgeman et al drew similar conclusions where the mass
of quantity and quality.7,53 The type and amount of gas released yield (dry ash-free) was 55.1, 61.5, and 72.0% for wheat straw, reed
as off-gas ­during torrefaction depend on the raw material type canary grass, and willow, respectively, at 290°C for 30 min.12
(i.e., anatomical, molecular, and chemical composition) and tor-
refaction process conditions, including the process temperature Condensable products
and residence time. Water is the main condensable product of torrefaction. It is
released during drying when moisture evaporates and during dehy-
Mass and energy yield dration reactions between organic molecules. Acetic acid is also
A typical mass-and-energy distribution is shown in Table 3, a condensable torrefaction product that mainly originates from
which illustrates the preservation of mass and energy in the solid ­acetoxy- and methoxy-groups present as side chains in xylose units
product. One can see that besides the solids, mostly water is pro-
duced during torrefaction, and the energy content of the volatiles is
mainly preserved in the lipids and organics. Torrefaction operating ■ Gaseous product ■ Liquid product ■ Solid product
conditions and biomass properties have a significant impact on the
Willow Larch Straw
amount of solid residue remaining and the volatile and gaseous 100
products produced during torrefaction. Thus, solid, volatiles, and 90
Product yield (wt%)

gases analysis and results from different studies are not exactly 80
the same. 70
60
Prins et al conducted several torrefaction mass balances experi-
50
ments.14 This data is reproduced in Figure 7, which shows a large
40
difference in product distribution for different types of biomass. 30
Larch has the largest yield of solid product and smallest gas and 20
liquid yield, straw has higher gas and liquid fractions than woody 10
biomass, and willow is between woody biomass and straw. It was 0
n) n) n) n) n) n) n) n) n) n)
also found that an increase in torrefaction temperature leads to a
0 mi 0 mi 0 mi 0 mi 5 mi 0 mi 0 mi 0 mi 5 mi 0 mi
decrease in solid char yield and an increase in the volatile fractions, (5 10 (3 (6 (1 (3 (5 (3 (1 (3
0 °C °C ( 0°C 0°C 0°C 0°C 0°C 0°C 0°C 0°C
including liquid and noncondensable gases, which is a result of the 23 230 25 25 27 27 23 25 27 25
competition between charring and devolatilization reactions that
become more reactive at higher temperatures. Figure 7. Overall mass balance of several torrefaction experiments14

Table 3. Mass and energy distribution of willow at 280°C


for 17.5 min reaction time31
Condensable volatile yield (wt%)

14
Energy yield ■ Acetic acid ■ Lactic acid
12 ■ Water ■ Furfural
Reaction products Mass yield (lower heating value ■ Formic acid ■ Hydroxyacetone
10 ■ Methanol ■ Phenol
[LHV], daf) (%)
8
Solid 87.5 94.9
6
Lipids 1.4 3.4
4
Organics 1.7 1.6
2
Gases 1.4 0.1
0
Water 8.0 0.0 230°C 250°C 270°C 280°C 300°C
(50 min) (30 min) (15 min) (10 min) (10 min)
Note: (daf: dry and ash-free basis)
Figure 8. Condensable volatiles yields during torrefaction14

© mary ann liebert, inc. • Vol. 7 No. 5 • october 2011 INdustrial Biotechnology 391
research

(part of the hemicellulose fraction). Prins et al proved that smaller CO2 production.11,14,52 In addition, a kinetic study on the generation
quantities of formic acid, lactic acid, furfural, hydroxyl acetone, of noncondensable gases shows that the gases are formed through
and traces of phenol are also present in the volatile component.14 parallel independent first-order reactions.13
Yields of most of the condensable volatiles will increase with the The composition of the noncondensable volatiles produced dur-
torrefaction temperature, as is shown in Figure 8. As a result, more ing different torrefaction conditions is shown in Figure 10. The
energy is transferred to the volatiles in the form of combustibles, ratio of CO to CO2 increased with temperature because cellulose and
such as methanol and acetic acid. lignin decompose at higher temperatures.14
Figure 11 shows the gas composition of the noncondensable
Noncondensable products products over time, which do not total 100% because only the main
The major gases formed during torrefaction are carbon diox- components (CO2 and CO) are shown.14 These results were obtained
ide and carbon monoxide with traces of other gases, as is shown by torrefaction of larch and willow at 250°C for 5, 15, and 30 min.
in Figure 9. 10,54 Carbon monoxide is the main source of the It was found that the ratio of CO2 to CO decreases with time, in line
noncondensable CV of torrefaction products. with the theory that CO is formed in a secondary reaction.14
The formation of CO2 may be explained by decarboxylation of
acid groups in wood and other herbaceous biomass. The formation Torrefaction process efficiency
of CO cannot be explained by dehydration or decarboxylation reac- The torrefaction process increases the energy density and
tions. The increased CO formation is reported to be caused by the improves the overall fuel properties of the torrefied product.
reaction of carbon dioxide and steam with porous char.55 Traces However, torrefaction is an energy-intensive pretreatment operation
of hydrogen and methane are also detected in noncondensable due to the high temperatures used in the process. Reusing the excess
products. A gas composition comparison between wood and agri- heat generated in the process will have a significant impact on the
cultural residues indicates that the latter is characterized by higher overall efficiency of the system. Also, both the condensable and
noncondensable volatiles have a significant impact on the energy
100 efficiency of the process. To achieve high energy efficiency at low
■ Mass composition cost, innovative torrefaction technologies have been developed to
■ Energy composition
75 capture excess heat and volatile energy; the Energy Research Center
of the Netherlands’ (ECN’s) torrefaction process is one of the most
Percent

50 mature, as is shown in Figure 12.10


The depicted process layout is based on the direct heating of
biomass during torrefaction by means of recycled hot gas. The
25
hot gas consists of the torrefaction gas and is re-pressurized to
compensate for the pressure drop in the recycle loop. It heats the
0 recycled gas to deliver the required heat demand in the torrefaction
CO CO2 Other
reactor. Combustion of the liberated torrefaction gas produces the
Figure 9. Main permanent gases compositions10,54 necessary heat for torrefaction and pre-drying. A utility fuel can be
used when the energy content of the torrefaction gas is insufficient

■ Carbon monoxide ■ Carbon dioxide


Noncondensable volatile yield (wt%)

3.5
Larch Straw Willow Larch
3.0
Composition of non-condensable

100
2.5 ■ Carbon monoxide 90
■ Carbon dioxide 80
gas phase (vol %)

2.0
70
1.5 60
50
1.0
40
0.5 30
0.0 20
230°C 250°C 270°C 290°C 250°C 10
(50 min) (30 min) (15 min) (10 min) (30 min) 0
5 min 15 min 30 min 5 min 15 min 30 min

Figure 11. Change in gas composition over time for torrefaction at


Figure 10. Noncondensable volatiles yields for larch and straw14 250°C 14

392 INdustrial Biotechnology october 2011


research

to thermally balance the torrefaction process and to provide stabil- content of about 10% wt, which makes up a total of 60% wt of
ity and control of the combustion process. Bergman et al identified incombustible components of the total emissions. The exact amount
this process concept as the most promising for torrefaction, which is determined by specific conditions and by the moisture content of
achieves autothermal operation when the total heat demand of the the biomass feed.
process (drying and torrefaction) is balanced by the energy content Bergman et al10,11 and Bergman32 further examined the CV of the
of the torrefaction gas.10 The torrefaction conditions (temperature torrefaction gas experimentally, while mass-and-­energy balance
and residence time) are the crucial variables to tune the thermal thermal-process efficiency, autothermal operation, and combustibil-
balance (i.e., the energy yield of torrefaction and hence the energy ity of the torrefaction gas were investigated by means of process sim-
content of the torrefaction gas). ulations. In their studies, the yield of reaction water varied between
The torrefaction gas consists of a wide variety of combustible 5 and 15% wt, resulting in a concentration of 50–80% wt in the tor-
organic components. However, the main components of torrefaction refaction gas (excluding free water from the feedstock). The reaction
gas are incombustible (water and CO2). The given product distribu- water yield increased with residence time and temperature, while its
tion in Figures 6–8 shows that the torrefaction gas is rather wet. concentration decreased. Consequently, the relative contribution of
Even when completely dry biomass is torrefied, the torrefaction combustible products increases with increased temperature and resi-
gas is expected to have a water content of over 50% wt and a CO2 dence time, as does the CV, which ranges from 5.3–16.2 ­MJ/­Nm3.
Despite the high water content of the torrefaction
gas, the CV is relatively high. It can be compared to
Air
Utility fuel producer gas from air-blown biomass gasification
Flue gas (4–7 MJ/Nm3) and to syngas from an indirectly ­heated
gasification process (15–20 MJ/Nm3). Based on this
Torrefaction Combustion comparison, the torrefaction gas should be combus-
gases Gas
Recycle Torrefied tible and can play an important role in the torrefac-
biomass tion process.10,11,32 Typical experimental results for
Biomass Drying Torrefaction Cooling torrefaction mass and energy yields and gas-phase
composition for willow are given in Figure 13.
DP
Solid torrefied biomass properties
Torrefaction of biomass significantly changes its
Flue gas Heat exchange physical and chemical properties like moisture con-
Flue gas
tent, density, grindability, pelletability, hydrophobic-
ity, CV, proximate and ultimate composition, and
Figure 12. The envisaged conceptual structure of ECN’s torrefaction process10
storage behaviors in terms of off-gassing, spontane-
ous combustion, and self heating.15

Gas phase composition Physical properties


CO = 0.1% Moisture content
CO2 = 3.3% Normally, feedstock moisture content ranges
H2O = 89.3% from 10–50%, but because torrefaction is a deep
Acetic acid = 4.8% drying process, moisture content is reduced to
Furfural = 0.2% 1–3% on a weight basis, depending on the torrefac-
Methanol = 1.2% tion conditions. 31 Typically, torrefaction achieves
Formic acid = 0.1%
an equilibrium moisture content of 3% and a
Remainder = 1.0%
Torrefaction ­reduction of mass by 20–30% (primarily by release
(32 min, 260°C) of water, carbon oxides, and volatiles), while
retaining 80–90% of the wood’s original energy
Feed: Willow Torrefied willow content.9 Reduction in moisture during torrefaction
Size: 10–30 mm Size: 10–30 mm provides three main benefits: (1) reduced moisture
LHV = 14.8 MJ/kg LHV = 18.5 MJ/kg level for the conversion process, (2) reduced trans-
Mass yield: 75.3%
MC = 14.4% (w.b.) MC = 1.9% (w.b.) portation costs associated with ­moving unwanted
Energy yield: 94.3%
Fixed carbon: 16.8% Fixed carbon: 22.1% water, and (3) the prevention of biomass decompo-
sition and moisture absorption during storage and
Figure 13. Typical experimental results of torrefaction of willow54 ­transportation.

© mary ann liebert, inc. • Vol. 7 No. 5 • october 2011 INdustrial Biotechnology 393
research

Table 4. Bulk density (dry basis) of Eucalyptus grandis wood which is mainly associated with the breakdown of the ­hemicellulose
in three different treatments56 matrix and depolymerization of the cellulose, ­resulting in decreased
fiber length.10,31 Particle length is also decreased, but not the ­diameter
Treatment Bulk density (g/cm3) Percentage loss
per se, resulting in better grindability, handling characteristics, and
Control* 0.85 a
— flowability through processing and transportation systems. Also
220°C 0.83 a
2.35 during the torrefaction process, the biomass tends to shrink, become
lightweight, flaky, and fragile, and lose its mechanical strength,
250°C 0.79b
7.06
making it easier to grind and pulverize.1
280°C 0.73 c
14.12 Bergman and Kiel conducted studies on the energy requirements
Note: Means followed by the same letter are statistically similar at the 5% for grinding raw and torrefied biomass like willow, woodcuttings,
probability level. demolition wood, and coal using a heavy-duty cutting mill.31 They
*Average moisture content of control treatment = 15%. concluded that power consumption reduces dramatically when bio-
mass is first torrefied. This reduction ranges from 70–90%, based on
the conditions under which the material is torrefied. Bergman and
Bulk and energy density Kiel also found that the capacity of the mill increases by a factor
Mass loss in the form of solids, liquids, and gases during of 7.5–15%.31 The most important observation they made was that
t­orrefaction cause the biomass to become more porous. This results the size-reduction characteristics of torrefied biomass were similar
in significantly reduced volumetric density, typically between to coal. Phanphanich and Sudhagar also observed a decrease in
180  and 300 kg/m 3, depending on initial biomass density and the grinding energy from 237.7–37.6 kWh/t, from raw to torrefied
­torrefaction conditions.31 Oliveira-Rodrigues and Rousset’s study on ­forest biomass at 300°C for 30 min.57 Figure 14 gives the typical
the effect of torrefaction on energy properties of Eucalyptus gran- grinding energy reduction for torrefied wood.58
dis wood indicated a bulk density loss of 14.12% when the wood
was torrefied at 280°C for 30 min.56 Table 4 shows the loss in bulk Particle size distribution, sphericity, and particle surface area
density for torrefaction temperatures ranging from 220–280°C.56 In Particle-size distribution curves, sphericity, and surface area are
spite of reduced bulk density, the energy density increases. Bergman important parameters for understanding flowability and combustion
et al observed that the calorific density of pine woodchips increased behavior during cofiring. Many researchers observed that ground,
from 11–20 MJ/kg.11 Many researchers have observed that the torrefied biomass produced narrower, more uniform particle sizes
energy density of both lignocellulosic and woody biomass increased compared to untreated biomass due to its brittle nature, which is
significantly after torrefaction (>60%). similar to coal. Phanphanich and Sudhagar studied torrefied pine
chips and logging residues and found that smaller particle sizes
Grindability are produced compared to untreated biomass.57 In addition, they
Biomass is highly fibrous and tenacious in nature; fibers form observed that the particle distribution curve was skewed towards
links between particles and make handling the raw ground samples smaller particle sizes with increased torrefaction temperatures.57
difficult. During torrefaction, the biomass loses its tenacious nature, Torrefaction also significantly influences the sphericity and
particle surface area. Phanphanich and Sudhagar’s results also
indicated that sphericity and particle surface area increased as the
900 torrefaction temperature was increased up to 300°C.57 For ground,
800 ■ 220°C torrefied chips, they found that the sphericity increased from 0.48%
Grinding energy (kWh/t)

700 ■ 240°C
■ 260°C to 0.62%, concluding that an increase in particle surface area or
600 decrease in particle size of torrefied biomass can be desirable prop-
500 erties for efficient cofiring and combustion applications.57 Also, the
400 bulk and particle densities of ground torrefied biomass increase
300 because the inter- and intraparticle voids generated after milling are
200 reduced.59–60 Research studies have indicated that ground torrefied
100 material results in a powder with favorable size distribution and
0 spherical particles, allowing torrefied powder to meet the smooth
Untreated 5 min 20 min 40 min 60 min
fluidization regime required for feeding it to entrained-flow pro-
Duration (min)
cesses (i.e., gasifier and pulverized coal).

Figure 14. Grinding energy of beech as a function of ­torrefaction Pelletability


duration, obtained with a Retsch ZM1 ultracentrifugal mill equipped Variability in feedstock quality due to differences in the types
with a 500‑mm grid58 of raw materials, tree species, climatic and seasonal variations,

394 INdustrial Biotechnology october 2011


research

Drying Torrefaction Size reduction Densification Cooling TOP pellets

Figure 15. Flow diagram for production of torrefied wood pellets32

1.8
storage conditions, and time significantly influence the quality of Willow C Untreated willow
Willow B (240°C:10 min)
biopellets.61 Torrefying the biomass before pelletization, however, 1.6 (240°C:60 min) Untreated
Willow A miscanthus
produces uniform feedstock with consistent quality. 1.4
(290°C:10 min)
Willow D Miscanthus C
Lignin in the biomass is considered to be the basic binding agent; (290°C:60 min) (240°C:10 min)

Atomic H: C ratio
1.2 Miscanthus B
(240°C:60 min)
thus, the pelletability of any biomass is evaluated based on lignin Miscanthus D
Miscanthus A
content. In general, higher amounts of lignin improve binding and 1.0 Bituminous
(290°C:60 min)
(290°C:10 min)
coal Lignite
reduce the severity of process conditions. The torrefaction process 0.8
opens more lignin‑active sites by breaking down the hemicellulose
0.6
matrix and forming fatty unsaturated structures, which creates Charcoal

better binding. Densification following torrefaction has been con- 0.4 Anthracite

sidered by several researchers.9,11,32,62–63 These studies indicate that, 0.2


compared to raw biomass pelletization, the required pressure and 0
energy consumption can be reduced by a factor of 2 when the mate- 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
rial is densified at a temperature of 225°C. In addition, densification Atomic O: C ratio
tests on torrefied biomass using a Pronto-Press indicated that the
density of torrefied and pelletized material varied in the range of Figure 16. Van Krevelen diagram for torrefied willow and
750–850 kg/m3 compared to conventional wood pellets, which are miscanthus64
in the range of 520–640 kg/m3. The energy density and CV have
also increased by approximately 70–80%, which is comparable to
sub-bituminous coal.31 Figure 15 shows the flow diagram for the and gasification processes. In torrefaction studies of reed canary
TOP process that Bergman proposed.32 grass and wheat straw torrefaction at 230, 250, 270, and 290°C for
30 min residence times, Bridgeman et al found that the moisture
Chemical compositional changes content decreases from an initial value of 4.7–0.8%.12 They found
Besides improving physical attributes, torrefaction also results in that carbon increased 48.6–54.3%, while the hydrogen and nitrogen
significant changes in proximate and ultimate composition of the content decreased from 6.8–6.1% and 0.3–0.1%, respectively. In
biomass and makes it more suitable for fuel applications. Sadaka another study, Bridgeman et al made a Van Krevelen diagram for
and Negi’s15 study on torrefaction of wheat straw, rice straw, and torrefied willow and miscanthus, as is shown in Figure 16.64 It is
cotton gin waste at 200, 260, and 315°C for 60, 120, and 180 min, clear that at higher temperatures and residence times, the atomic
respectively, concluded that moisture content was reduced at the O/C and H/C ratios are closer to that of lignite coal. Table 5 shows
conditions (260°C for 120 min) for all three feedstocks by 70.5, 49.4, the effect of different torrefaction temperatures on ultimate compo-
and 48.6%, and the heating value increased by 15.3, 16.9, and 6.3%, sitional changes in woody and herbaceous biomass.12 Table 6 shows
respectively. Zanzi et al, in their study on miscanthus torrefaction, the physiochemical composition of pine and heat-treated pine at
made similar observations, where increasing temperature from 230 torrefaction temperatures from 240–290°C and indicates that as tor-
to 280°C and time from 1–3 h increased the carbon content and refaction temperature increases, fixed carbon increases and volatiles
decreased the hydrogen, nitrogen, and oxygen content.7 At 280°C, and moisture content decrease.65
the carbon content increased to about 52% from an initial value of
43.5%, while hydrogen and nitrogen content decreased from 6.49 Calorific value
to 5.54% and 0.90 to 0.65%, respectively, for 2 h of torrefaction. Biomass loses relatively more oxygen and hydrogen than ­carbon
In general, increased torrefaction temperatures result in increased during torrefaction, which increases the calorific value of the prod-
carbon content and decreased hydrogen and oxygen content due to uct.6 The net CV of torrefied biomass is 18–23 MJ/kg (lower heating
the formation of water, CO, and CO2. value [LHV], dry) or 20–24 MJ/kg (higher heating value [HHV],
This process also causes the hydrogen-to-carbon (H/C) and dry).11,66 The mass and energy in the torrefied biomass is preserved
oxygen-to-carbon (O/C) ratios to decrease with increasing tor- in the solid product for a long time, as the material does not degrade
refaction temperature and time, which results in less smoke and with time. Table 5 provides the ultimate analysis and elemental
water-vapor formation and reduced energy loss during combustion composition of torrefied woody and herbaceous biomass proper-

© mary ann liebert, inc. • Vol. 7 No. 5 • october 2011 INdustrial Biotechnology 395
research

Table 5. Ultimate analysis, moisture content, and HHV (dry ash free basis) of untreated and torrefied biomass12

Raw Torrefaction temperature


Reed canary grass 0°K 503°K 523°K 543°K 563°K

C (%) 48.6 49.3 50.3 52.2 54.3

H (%) 6.8 6.5 6.3 6.0 6.1

N (%) 0.3 0.1 0.0 0.1 0.1

O (%) 37.3 N/A 37.0 37.3 36.3

Moisture (%) 4.7 2.5 1.9 1.3 1.2

CV (kJ/kg) 19,500 N/A 20 000 20 800 21 800

Wheat straw

C (%) 47.3 48.7 49.6 51.9 5.6

H (%) 6.8 6.3 6.1 5.9 1.0

N (%) 0.8 0.7 0.9 0.8 27.6

O (%) 37.7 N/A 35.6 33.2 0.8

Moisture (%) 4.1 1.5 0.9 0.3 0.8

CV (kJ/kg) 18,900 19 400 19 800 20 700 22 600

Willow

C (%) 49.9 50.7 51.7 53.4 54.7

H (%) 6.5 6.2 6.1 6.1 6.0

N (%) 0.2 0.2 0.2 0.2 0.1

O (%) 39.9 39.5 38.7 37.2 36.4

Moisture (%) 2.8 0.5 0.1 0.1 0.0

CV (kJ/kg) 20,000 20 600 20 600 21 400 21 900

Table 6. Physiochemical analysisa of pine and heat-treated pine65


T (°C)b pc 240 250 260 270 290

Fixed carbon, % 20.64 23.55 25.59 25.69 29.38 35.39

Elementary analysis
C, % 50.98 51.14 51.93 53.78 53.57 58.08
O, % 42.80 42.70 42.18 40.66 40.67 36.40

Pentosans, % 9.61 5.93 5.90 3.10 2.54 1.40

Lignin, % 22.84 24.90 28.72 33.44 39.23 53.47

Extractables, % d
14.67 8.19 14.09 19.35 16.49 17.98

Moisture, % e
10.80 5.66 4.08 3.96 3.76 3.88

Yield, % — 86.2 81.8 75.7 66.4 48.8

a
In each case, the mean result given was obtained from a minimum of four different experiments. bTorrefaction time: 30 min. cNative pine. dNeutral-solvent extractables
(ethanol, benzene, boiling water). ePowder samples left at the laboratory atmosphere still had a constant humidity.

396 INdustrial Biotechnology october 2011


research

ties as compared to raw materials. Sadaka and Negi also observed gravimetric measurement to determine the degree of water uptake.
that the highest heating value of 22.75 MJ/kg, or 9761 BTU/lb, was Bergman’s study indicated that raw pellets swelled rapidly and dis-
achievable at torrefaction conditions of 315°C and 3 h.15 integrated into original particles.32 Torrefied pellets produced under
optimal conditions, however, did not disintegrate and showed little
Storage aspects of torrefied biomass water uptake (7–20% on a mass basis). Bergman also concluded that
Off-gassing torrefaction conditions play a vital role in the hydrophobic nature
Storage issues like off-gassing and self-heating may also be of biomass.32 Sokhansanj et al compared the moisture uptake of
insignificant in torrefied biomass, as most of the solid, liquid, and the torrefied biomass to the untreated biomass and found a 25%
gaseous products that are chemically and microbiologically active decrease in water uptake when compared to the control, as shown
are removed during the torrefaction process. Kuang et al’s67 and in Figure 17.71
Tumuluru et al’s 68 studies on wood pellets concluded that high stor- Biomass is porous, often moist, and prone to off-gassing and
age temperatures of 50°C can result in high CO and CO2 emissions, self-heating due to chemical oxidation and microbiological activity.
and the concentrations of these off-gases can reach up to 1.5% and In general, the biomass moisture content plays an important role in
6% for a 60-day storage period. These emissions were also found to initiating chemical and microbial reactions. Moisture content cou-
be sensitive to relative humidity and product moisture content. The pled with high storage temperatures can cause severe off-­gassing
same researchers at the University of British Columbia conducted and self-heating from biomass-based fuels. Another important stor-
studies on off-gassing from torrefied woodchips and indicated that age issue for ground torrefied biomass is its reactivity in powder
CO and CO2 emissions were very low — nearly one-third that of form, which can result in fire during storage. It is preferred to store
regular woodchips at room temperature (20°C). This could be due the torrefied biomass in an inert environment to avoid incidents of
to low moisture content and reduced volatile content, which could spontaneous combustion. In his laboratory-scale combustion stud-
result in less reactivity with the storage environment. ies of torrefied wood, Kiel found that it is highly reactive, similar
to coal.53
Hydrophobicity Table 7 shows the product characteristics of torrefied biomass
In general, the uptake of water by raw biomass is due to the pres- compared to coal and wood, revealing that torrefied wood closely
ence of OH groups. Torrefaction produces a hydrophobic product by resembles charcoal. The major difference between charcoal and
destroying OH groups and causing the biomass to lose the capacity ­torrefied wood is the volatile content. Volatiles are lost during
to form hydrogen bonds.69 Due to these chemical rearrangement charcoal production, which also means a possible loss of energy.72
reactions, nonpolar unsaturated structures are formed, which pre- On the other hand, during torrefaction, most of the volatiles are
serve the biomass for a long time without biological degradation, retained. It is also recommended that every form of carbonization
similar to coal.31,70 be avoided during torrefaction. As Table 7 shows, torrefied pellets
Bergman determined the hydrophobicity of torrefied pellets by have product characteristics, like handling, milling, and transport
immersing them in water for 15 h.32 Their hydrophobic nature was requirements, similar to coal.65 Torrefied pellets allow for higher
evaluated based on the state of the pellet after this period and by cofiring percentages up to 40% due to matching fuel properties
with coal, and they can use the existing equipment setup for coal.
20
Torrefied biomass applications
Pulverized fuel combustion in coal-fired power stations and
16 25%
Moisture uptake (% mass)

decrease entrained-flow gasification are particularly interesting product


outlets for biomass. In both applications, biomass has to be fed
12 to the reactor as a powder, which is difficult, costly, and achiev-
able only at very low capacity in classical coal mills. Due to this
limitation, wood pellets are currently the state-of-the-art for
8
cofiring, as they consist of sufficiently small particles. Consequently,
◆ Control
wood pellets also have some limitations in terms of energy content
4 ■ 250°C and moisture uptake during storage and transportation. Torrefied
▲ 300°C
biomass, because it is energy-dense and hydrophobic in nature, can
0 be a good replacement for wood pellets in cofiring and gasification
0 20 40 0 plants. The high fuel quality of torrefied biomass makes it very
Exposure time (h) attractive for combustion and gasification applications. Due to high
calorific values, the thermal energies of the combustion and gas-
Figure 17. Moisture uptake by the torrefied wood pellets made from ification system can be improved significantly.31 However, data is
a 0.8 mm particle size71 lacking on milling, handling, storing, transporting, and combusting

© mary ann liebert, inc. • Vol. 7 No. 5 • october 2011 INdustrial Biotechnology 397
research

Table 7. Indicative properties of different biomass and coal-based fuels72


Wood Wood pellets Torrefied pellets Charcoal Coal

Moisture content (% wt) 30–45 7–10 1–5 1–5 10–15

Calorific value (MJ/kg) 9–12 15–16 20–24 30–32 23–28

Volatiles (% db) 70–75 70–75 55–65 10–12 15–30

Fixed carbon (% bd) 20–25 20–25 28–35 85–87 50–55

Bulk density (kg/l) 0.2–0.25 0.55–0.75 0.75–0.85 ~0.2 0.8–0.85

Volumetric energy density (GJ/m3) 2.0–3.0 7.5–10.4 15.0–18.7 6.0–6.4 18.4–23.8

Dust Average Limited Limited High Limited

Hygroscopic properties Hydrophilic Hydrophilic Hydrophobic Hydrophobic Hydrophobic

Biological degradation Yes Yes No No No

Milling requirement Special Special Classic Classic Classic

Handling requirements Special Easy Easy Easy Easy

Product consistency Limited High High High High

Transport cost High Average Low Average Low

large amounts of torrefied biomass at a commercial scale. Common 3. T orrefaction of the biomass helps in developing a uniform
torrefied biomass applications include: (1) high-quality smokeless feedstock with minimum variability in moisture content.
solid fuels for industrial, commercial, and domestic applications, 4. Torrefaction of biomass improves (a) energy density, grindability,
(2) solid fuel for cofiring directly with pulverized coal at electric and pelletability index ratings, (b) ultimate and proximate com-
power plants, (3) an upgraded feedstock for fuel pellets, briquettes, position by increasing the carbon content and CV and decreas-
and other densified biomass fuels, and (4) high-quality biomass ing the moisture and oxygen content, and (c) biochemical com-
solid fuel for advanced bioenergy application. position by decomposing the hemicelluloses and softening the
lignin, which results in better binding during pelletization.
Conclusions 5. Ground torrefied biomass has improved sphericity, particle
Interest in research on torrefaction of biomass materials is grow- surface area, and particle size distribution.
ing. Its potential to improve the quality of both herbaceous and 6. Densification of torrefied material reduces specific energy
woody materials provides a path for using these resources in consumption and increases throughput by about a factor of 2
many energy applications. This review has provided insight into compared to raw or untreated biomass.
the ­limitations of raw biomass materials to meet the needs of 7. During torrefaction the biomass loses most of the low energy
energy providers and has discussed options for improving biomass content of the material, like (a) solids, which include original
­resources for use in power and liquid-fuel production. Based on this sugar structures and other newly formed polymeric structures,
review, nine areas of discussion are summarized below. These areas and (b) liquids, which include condensables, like water, organ-
provide a brief understanding of the potential use of torrefaction as ics, and lipids, and (c) gases, which include H2, CO, CO2 and
a means of improving the physical, chemical, and rheological char- CH4, CxHy, toluene, and benzene.
acteristics of biomass materials. As with most biomass-upgrading 8. Torrefaction preserves the biomass for a long time without
options, however, torrefaction, and the subsequent densification biological degradation due to the chemical rearrangement
processes, need to be evaluated for energy inputs to understand the reactions and formation of nonpolar unsaturated structures.
logistical advantages and end use of the product. 9. Torrefied biomass has been successfully used as an upgraded
1. Torrefied biomass, in general, defines a group of products solid fuel in electric power plants and gasification plants.
resulting from the partially controlled and isothermal pyrolysis Not all aspects of torrefaction and its influence on other pro-
of biomass occurring at the 200–300°C temperature range. cessing operations have been explored. Opportunities for future
2. The most common torrefaction reactions include (a) devolatil- research can include:
ization and carbonization of hemicelluloses, and (b) depoly- a. Optimizing torrefaction processes based on activation energies
merization and devolatilization of lignin and cellulose. required to degrade the cellulose, hemicelluloses, and lignin.

398 INdustrial Biotechnology october 2011


research

b. U  nderstanding the torrefaction process at a molecular level by REFERENCES


identifying different functional groups and energies associated 1. Arias BR, Pevida CG, Fermoso JD, Plaza MG, Rubiera FG, and Pis-Martinez JJ.
with breaking the bonds. Influence of torrefaction on the grindability and reactivity of woody biomass.
c. Understanding the spontaneous torrefaction process reactions Fuel Process Technol 89(2), 169–175 (2008).
using Fourier transform infrared (FTIR) and Raman spectros- 2. United Nations (UN). Kyoto protocol to the United Nations framework conven-
tion on climate change [Online]. Available at: http://unfccc.int/resource/docs/
copy. convkp/kpeng.pdf [August 3, 2011] (1998).
d. Understanding the severity of the torrefaction process based on
3. Evergreen Renewables, LLC. Biomass torrefaction as a preprocessing step
color changes using the Hunter colorimeter. for thermal conversion: Reducing costs in the biomass supply chain [Online].
e. Studies on thermogravimetrics to establish the weight-loss Available at: http://evergreenrenewable.com/welcome_files/Biomass%20torre-
kinetics. faction.pdf [August 3, 2011] (2009).
f. Studies on microstructural changes in biomass at different 4. Pipatmanomai S. Overview and experiences of biomass fludized bed gasification
temperature regimes. in Thailand. J Sustainable Energy Environ Special Issue, 29-11 (2011).
g. Testing integrated processes such as torrefaction and densifi- 5. Sierra R, Smith A, Granda C, and Holtzapple MT. Producing fuels and chemi-
cals from lignocellulosic biomass. Chemical Engineering Process, Society of
cation. Biological Engineering Special Section: Biofuels 104, S10–S18 [Online]. Available
h. Calculating energies associated with condensable and noncon- at: http://www.aiche.org/uploadedFiles/SBE/MemberCenter/0808S10(2).pdf
densable products produced during torrefaction and the ability [August 3, 2011] (2008).
to reuse them to improve the overall process efficiency. 6. Uslu A, Faaij A, and Bergman PCA. Pre-treatment technologies, and their effect
i. Understanding the storage behavior of torrefied biomass in on international bioenergy supply chain logistics: Techno-economic evalua-
tion of torrefaction, fast pyrolysis, and pelletisation. Energy 33(8), 1206–1223
terms of off-gassing and spontaneous combustion at different (2008).
storage temperatures.
7. Zanzi R, Ferro DT, Torres A, Soler PB, and Bjornbom E. Biomass torrefaction.
In: The 6th Asia-Pacific International Symposium on Combustion and Energy
Utilization. Kuala Lumpur, Malaysia, 20–22 May 2002 (2002).
AC K N OW L E D G M E N T S 8. Bourgeois JP and Doat J. Torrefied wood from temperate and tropical species:
The authors acknowledge Leslie Park Ovard for her valu- Advantages and prospects. In: Bioenergy 84, 3, 153–159. Egneus H, Ellegard
A, Okeefe P, and Kristofferson L (eds), Elsevier Applied Science, London, United
able contribution to the manuscript, and Gordon Holt, Allen Kingdom. (1984).
Haroldsen, Lisa Plaster, and Huan-Chen Tang, from Idaho National
9. Lipinsky ES, Arcate JR, and Reed TB. Enhanced wood fuels via torrefaction.
Laboratory’s R&D Publications Support Team for their editorial Fuel Chemistry Division Preprints, 47(1), 408–410 (2002).
and graphics creation assistance. This work is supported by the
10. Bergman PCA, Boersma AR, Zwart RWH, and Kiel JHA. Torrefaction for biomass
US Department of Energy, under DOE Idaho Operations Office co-firing in existing coal-fired power stations. Report ECN-C--05-013, ECN,
Contract DE-AC07-05ID14517. Accordingly, the US Government Petten (2005a).
retains and the publisher, by accepting the article for publication, 11. Bergman PCA, Boersma AR, Kiel JHA, Prins MJ, Ptasinski KJ, and Janssen FGGJ.
acknowledges that the US Government retains a nonexclusive, Torrefied biomass for entrained-flow gasification of biomass. Report ECN-C--
05-026, ECN, Petten (2005b).
paid-up, irrevocable, worldwide license to publish or reproduce
the published form of this manuscript, or allow others to do so, 12. Bridgeman TG, Jones JM, Shield I, and Williams PT. Torrefaction of reed canary
grass, wheat straw, and willow to enhance solid fuel qualities and combustion
for US Government purposes. properties. Fuel 87(6), 844–856 (2008).
13. Prins MJ, Ptasinski KJ, and Janssen FGGJ. Torrefaction of wood: Part 1, Weight
loss kinetics. J Anal Appl Pyrolysis 77(1), 28–34 (2006a).
US DEPARTMENT OF ENERGY DISCLAIMER
14. Prins MJ, Ptasinski KJ, and Janssen FGGJ. Torrefaction of wood: Part 2, Analysis
This information was prepared as an account of work spon- of products. J Anal Appl Pyrolysis 77(1), 35–40 (2006b).
sored by an agency of the US government. Neither the US govern- 15. Sadaka S and Negi S. Improvements of biomass physical and thermochemical
ment nor any agency thereof, nor any of their employees, makes characteristics via torrefaction process. Environmental Progress Sustainable
any warranty, express or implied, or assumes any legal liability Energy 28(3), 427–434. doi: 10.1002/ep.10392 (2009).
or responsibility for the accuracy, completeness, or usefulness 16. Moire L, Schmutz A, Buchala A, Yan B, Stark RE, and Ryser U. Glycerol is a
suberin monomer: New experimental evidence for an old hypothesis. Plant
of any information, apparatus, product, or process disclosed, or Physiol 1999 March, 119(3), 1137–1146 (1999).
represents that its use would not infringe privately owned rights.
17. Esau K. Plant Anatomy, 2nd ed. John Wiley & Sons, New York, New York (1964).
Any reference herein to any specific commercial product, process,
18. Mohan D, Pittman CU, and Steele PH. Pyrolysis of wood/biomass for bio-oil:
or service by trade name, trademark, manufacturer, or otherwise,
A critical review. Energ Fuel 20(3), 848–889 (2006).
does not necessarily constitute or imply its endorsement, recom-
19. Shafizadeh F. Pyrolytic reactions and products of biomass. In: Fundamentals
mendation, or favoring by the US Government or any agency of Biomass Thermochemical Conversion. 183–217. Overend RP, Milne TA, and
thereof. The views and opinions of the authors expressed herein Mudge LK (eds), Elsevier Applied Science, London, United Kingdom (1985).
do not necessarily state or reflect those of the US Government or 20. Williams PT and Besler S. The influence of temperature and heating rate on the
any agency thereof. slow pyrolysis of biomass. Renewable Energy 7(3), 233–250 (1996).

© mary ann liebert, inc. • Vol. 7 No. 5 • october 2011 INdustrial Biotechnology 399
research

21. Stamm AJ. Thermal degradation of wood and cellulose. Ind Eng Chem 48, ing. Wood Sci Technol 34(6), 473–480 (2000).
413–417 (1956).
44. Kauman WG. Contribution to the theory of cell collapse in wood: investigations
22. Esteves B, Marques AV, Domingos I, and Pereira H. Influence of steam heating with eucalyptus regnans. Maderas. Ciencia y tecnología 4(1), 77–99 (2002).
on the properties of pine (Pinus pinaster) and eucalypt (Eucalyptus globulus)
wood. Wood Sci Technol 41(3), 193–207 (2007). 45. Tiemann HD. Eucalyptus lumber. (Abstr. Rep. For. Prod. Lab., Madison,
Wisconsin) Hardwood Rec Sept. 25, Oct. 10, 1913.
23. Borrega M and Kärenlampi PP. Effect of relative humidity on thermal degrada-
tion of Norway spruce (Picea abies) wood. J Wood Sci 54, 323–328 (2008). 46. Greenhill WL. Collapse and its removal: some recent investigations with euca-
lyptus regnans. Coun Sci Industr Res (Aust.) Pamph. 75 (1938).
24. Sterling C. Effect of moisture and high temperature on cell walls in plant tis-
sues. Food Res 20(5), 474–479 (1955). 47. Ellwood EL. The seasoning of rotary peeled veneer from Eucalyptus regnans
F.v.M. Aust J Appl Sci 3(1), 53–70 (1952).
25. Kärenlampi, PP, Tynjala P, and Strom P. Phase transformation of wood cell wall
water. J Wood Sci 51, 118–123 (2005). 48. Ellwood EL, Gottstein, JW, and Kauman WG. A laboratory study of the vapour
drying process: Part III. Vapour drying of timber in joinery and railway sleeper
26. Björklund Jansson M. Two methods of extraction and GC-analysis of lipophilic sizes. CSIRO (Aust.), Div. For. Prod. Technol. Paper No. 14 (1961).
wood extractives. Report #3, Nordic Standardization Program (NSP) (2008).
49. Demirbas A. Pyrolysis mechanisms of biomass materials. Energy Sources Part A
27. Fernández de Simón B, García-Vallejo MC, Cadahía E, Arrabal C, and Cortijo M. 31(13), 1186–1193. doi: 10.1080/15567030801952268 (2009).
Analysis of lipophilic compounds in needles of Pinus pinea L. Ann For Sci 58,
449–454 (2001). 50. Bourgeois J and Guyonnet R. Characterization and analysis of torrefied wood.
Wood Sci Technol 22(2), 143–155 (1988).
28. GuiJun W, YongHao L, Jian D, JiangHong K, and YunLiang Z. Pretreatment of
biomass by torrefaction. Chin Sci Bull 56(14), 1442–1448 (2011). 51. Evans RJ, Milne TA, and Soltys MN. Direct mass-spectrometric studies of the
pyrolysis of carbonaceous fuels: Part III. Primary pyrolysis of lignin. J Anal Appl
29. Kojiro K, Miki T, Sugimoto H, and Kanayama K. Destabilization of wood micro- Pyrol 9(3), 207–236 (1986).
structure caused by drying. Wood Material Science & Engineering 6(1–2), 69–74
doi:10.1080/17480272.2010.551545 (2011). 52. Deng J, Wang GJ, Kuang JH, Zhang YL, and Luo YH. Pretreatment of agricultural resi-
dues for co-gasification via torrefaction. J Anal Appl Pyrol 86(2), 331–337 )2009).
30. Thuvander F, Kifetew G, and Berglund LA. Modeling of cell wall drying stresses
in wood. Wood Sci Technol 36(3), 241–254 (2002). 53. Kiel JHA. ECN BO2-technology for biomass upgrading. BUS final meeting.
Wageningen, The Netherlands, 20 November 2007 (2007a).
31. Bergman PCA and Kiel JHA. Torrefaction for biomass upgrading. In: Proceedings
of the 14th European Biomass Conference & Exhibition. Paris, France, 17–21 54. Kiel JHA. Torrefaction for biomass upgrading into commodity fuels. In:
October 2005. Proceedings of the IEA Bioenergy Task 32 Workshop on Fuel Storage, Handling
and Preparation and System Analysis for Biomass Combustion Technologies.
32. Bergman PCA. Combined torrefaction and pelletization: The TOP process.
Berlin, Germany, 7 May 2007 (2007b).
Report ECN-C-05-073, ECN, Petten (2005).
33. Lam PK, Sokhansanj S, Bi T, and Lim CJ. Quantifying color of steam treated pel- 55. White RH and Dietenberger MA. Wood products: Thermal degradation and fire.
lets made from western Douglas Fir (Pseudotsuga Menziesii, L.). Trans ASABE In: The Encyclopedia of Materials: Science and Technology. Buschow KHJ, Cahn
(under review, 2011). RW, Flemings MC, Ilschner B, Kramer EJ, Mahajan S, and Veyssière P (eds.),
Elsevier Applied Science, Amsterdam, The Netherlands (2001).
34. Summa CA, de la Calle B, Brohee M, Stadler RH, and Anklam E. Impact of the
roasting degree of coffee on the in vitro radical scavenging capacity and con- 56. Oliveira-Rodrigues T and Rousset PLA. Effects of torrefaction on energy proper-
tent of acrylamide. LWT, 40, 1849-1854 (2007). ties of eucalyptus grandis wood. Cerne 15(4), 446–452 (2009).

35. Shafizadeh F and McGinnis GD. Chemical composition and thermal analysis of 57. Phanphanich M and Sudhagar M. Impact of torrefaction on the grindability and
cottonwood. Carbohyd Res 16, 273–277 (1971). fuel characteristics of forest biomass. Bioresource Technol 102(22), 1246–1253
(2011).
36. Terziev N. Effect of high-temperature and microwave treatment on microstruc-
ture of softwoods. In: Proceedings of the 4th COST E15 Workshop: Methods for 58. Repellin V, Govin A, Rolland M, and Guyonnet R. Energy requirement for fine
Improving Drying Quality of Wood. Santiago de Compostela, Spain, 30–31 May grinding of torrefied wood. Biomass Bioenerg 34(7), 923–930 (2010).
2002. 59. Mani S, Tabil LG, and Sokhansanj S. Grinding performance and physical prop-
37. Borrega M and Kärenlampi PP. Cell wall porosity in Norway spruce wood as erties of wheat and barley straws, corn stover, and switchgrass. Biomass
affected by high temperature drying. Wood Fiber Sci 43(2), 206–214 (2011). Bioenergy 27(4), 339–352 (2004).
38. Borrega M and Kärenlampi PP. Three mechanisms affecting the mechanical prop- 60. Esteban LS and Carrasco JE. Evaluation of different strategies for pulverization
erties of spruce wood dried at high temperatures. J Wood Sci 56, 87–94 (2010). of forest biomasses. Powder Technol 166(3), 139–151 (2006).
39. Weise U, Maloney T, and Paulapuro H. Quantification of water in different 61. Lehtikangas P. Quality properties of fuel pellets from forest biomass. Licentiate
states of interaction with wood pulp fiber. Cellulose 3, 189–202 (1996). Thesis, Report #4. Department of Forest Management and Products, University
of Agricultural Sciences. Uppsala, Sweden (1999).
40. Park S, Venditti RA, Jameel H, and Pawlak JJ. Changes in pore size distribution
during the drying of cellulose fibers as measured by differential scanning calo- 62. Reed TB and Bryant B. Densified biomass: A new form of solid fuel. Solar
rimetry. Carbohyd Polym 66(1), 97–103 (2006). Energy Research Institute Report #SERI–35, US Department of Energy, Division
of Solar Technology, Golden, CO (1978).
41. Van den Akker JA. Some theoretical considerations on the mechanical proper-
ties of fibrous structures. In: Formation and Structure of Paper. 205–241. F. 63. Koukios EG. Progress in thermochemical, solid state refining of biofuels: From
Bolam (ed.) Transactions of the Fundamental Research Symposium, Oxford, research to commercialization. Advanced Thermochem Biomass Conversion, 2,
United Kingdom, September 1961. Technical Section of the British Paper and (1993).
Board’s Makers Association, London, United Kingdom (1962).
64. Bridgeman TG, Darvell LI, Gudka B, Fisher EM, Jones JM, Waldron D, and
42. Kifetew G, Thuvander F, Berglund L, and Lindberg H. The effect of drying on Williams A. Milling and combustion properties of torrefied biomass: A potential
wood fracture surfaces from specimens loaded in wet condition. Wood Sci co-firing fuel. Workshop on Cofiring Biomass with Coal, Session 4: Torrefaction.
Technol 32(2), 83–94 (1998). Drax Power Station, United Kingdom, 25–26 January 2011.
43. Thuvander F, Wallström L, Berglund LA, and Lindberg KAH. Effects of an 65. Bourgeois J, Bartholin MC, and Guyonnet R. Thermal treatment of wood: analy-
impregnation procedure for prevention of wood cell wall damage due to dry- sis of the obtained product. Wood Sci Technol 23, 303–310 (1989).

400 INdustrial Biotechnology october 2011


research

66. Prins MJ. Thermodynamic analysis of biomass gasification and torrefaction. 70. Wooten JB, Crosby B, and Hajaligol MR. Evaluation of cellulose char structure moni-
Thesis submitted to Eindhoven. Technische Universiteit Eindhoven (2005). tored by 13C CP MAS NMR. Fuel Chem Division Preprints 46, 191–193 (2000).
67. Kuang X, Tumuluru JS, Bi XT, Lim CJ, Sokhansanj S, and Melin S. Rate and peak 71. Sokhansanj S, Peng J, Lim J, Bi X, Wang L, Lam P, Hoi J, Melin S, Tumuluru J,
concentrations of off-gas emissions in stored wood pellets — sensitivities to and  Wright C. Optimum torrefaction and pelletization of biomass feedstock.
temperature, relative humidity, and headspace volume. Ann Occup Hyg 53(8), TCS 2010 Symposium on Thermal and Catalytic Sciences for Biofuels and
789–796 (2009). Biobased Products. Iowa State University, Ames, Iowa, 21–23 September 2010.
68. Tumuluru JS, Kuang X, Sokhansanj S, Lim CJ, Bi X, and Melin S. Development of lab- 72. Kleinschmidt CP. Overview of international developments on torrefaction.
oratory studies on the off-gassing of wood pellets. Can Bio Eng 52, 8.1–8.9 (2010). Central European Biomass Conference 2011, Graz, Austria, January 26–28, 2011
69. Pastorova I, Arisz PW, and Boon JJ. Preservation of D-glucose oligosaccharides [Online]. Available at: http://www.bioenergytrade.org/downloads/grazklein-
in cellulose chars. Carbohyd Res 248, 151–165 (1993). schmidtpaper2011.pdf [August 3, 2011] (2011).

© mary ann liebert, inc. • Vol. 7 No. 5 • october 2011 INdustrial Biotechnology 401

You might also like